Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Multiple Proteases to Localize Oxidation Sites

  • Liqing Gu,

    Affiliation Department of Chemistry, University of Pittsburgh, Pittsburgh, PA, 15260, United States of America

  • Renã A. S. Robinson

    rena@pitt.edu

    Affiliation Department of Chemistry, University of Pittsburgh, Pittsburgh, PA, 15260, United States of America

Abstract

Proteins present in cellular environments with high levels of reactive oxygen and nitrogen species and/or low levels of antioxidants are highly susceptible to oxidative post-translational modification (PTM). Irreversible oxidative PTMs can generate a complex distribution of modified protein molecules, recently termed as proteoforms. Using ubiquitin as a model system, we mapped oxidative modification sites using trypsin, Lys-C, and Glu-C peptides. Several M+16 Da proteoforms were detected as well as proteoforms that include other previously unidentified oxidative modifications. This work highlights the use of multiple protease digestions to give insights to the complexity of oxidative modifications possible in bottom-up analyses.

Introduction

Reactive oxygen and nitrogen species in cellular environments can result in macromolecular damage [1]. Protein tertiary structure can be modified as a result of oxidative attack and lead to loss or gain of protein function [2]. Site-specific modifications can include the incorporation of oxygen atoms to amino acid side chains (e.g., methionine, phenylalanine, tyrosine, leucine, histidine), conversion to aldehydes, ketones, or carboxylic acids, addition of nitrogen containing groups (e.g., R-NO, R-NO2), formation of advanced glycation end products (e.g. imidazolinones, pyrralines), and introduction of lipid peroxidation products (e.g., 4-hydroxy-trans-nonenal, malondialdehyde) [35]. Key to understanding the events that affect protein function is the ability to characterize the distribution of oxidized proteoforms.

Techniques for the identification of proteoforms have been recently discussed [610]. Proteoforms can include molecules that arise due to the same post-translational modification (PTM) occurring at different amino acid residue positions in the protein. For example, a protein that incorporates a single oxygen atom during a free radical attack from hydrogen peroxide (H2O2) or superoxide anion, may exist in multiple locations. One population of the protein molecules can incorporate the oxygen at residue “A”, others incorporate at residue “B”, while the remaining molecules incorporate at both “A” and “B”. For the molecules with only a single oxygen addition, mass spectrometry (MS) measurements of intact protein would only detect a single M+16 Da species. Liquid chromatography (LC) or electrophoresis separations may be able to resolve the two proteoforms (i.e., A and B), however multiple dissociation methods such as collision activation dissociation (CAD) [11], infrared multiphoton dissociation [12], electron capture dissociation (ECD) [13], or electron transfer dissocation (ETD) [14] are necessary to localize the modification site.

Top-down and bottom-up protein analysis provides complementary information regarding protein sequence and PTMs [8,1518], especially for identification of oxidation sites [1921]. Top-down MS has been employed to characterize four oxidation sites in viral prolyl-r-hydroxylase [20] and for the identification of 250 isoforms of oxidized calmodulin [21]. Bottom-up proteomics is very useful for the verification of PTM types and sites although it can be challenging to identify the specific proteoform from which peptides originated. This is because shotgun analysis of all proteins extracted will lead to many similar peptides produced from various proteoforms. Because fragmentation of peptides is very accessible with CID and other dissociation methods, as compared to intact proteins, it is very practical to use bottom-up analyses to localize sites of oxidative modification. However, for complex biological samples the number of modification sites that can be characterized without extensive enrichment or separation strategies is generally low [22,23]. Therefore, it will be necessary to incorporate enrichment with our strategy of multiple proteases and iterative database searching to gain localize oxidative modification sites and obtain insight to the complexity of oxidized proteoforms present in complex mixtures.

Ubiquitin is a low molecular weight protein which has significant roles in protein turnover and degradation through its molecular chaperoning activity in the proteasome [24]. Ubiquitin is implicated in oxidative stress and disease [25,26]. The 76 amino acid sequence of this protein is highly conserved amongst eukaryotes, such as bovine and human [27]. Herein, we aimed to characterize the heterogeneity of proteoforms of ubiquitin using moderate oxidizing conditions [28] and bottom-up MS with multiple proteases. Chemical oxidizing conditions using Fenton chemistry [Fe(II)/H2O2] [28] rely on the metal serving as an electron donor to catalyze the formation of highly reactive hydroxyl radical (·OH) which can result in modification of amino acid side chains [5,29]. Previous studies have investigated oxidized forms of ubiquitin after exposure to peroxynitrite [30], electrochemical oxidation [31,32], and photochemical reactions [33] for the purpose of structural footprinting. The influence of N-terminal oxidation of Methionine (hereafter referred to as Met1-Ox) on protein structures and stabilities were examined by ion mobility spectrometry-mass spectrometry (IMS-MS)[34] and indicate that oxidation of Met1 can lead to destabilization of the native state and result in unfolded structures. Thus simple oxidized proteoforms can have a huge influence on protein structure.

Bottom-up LC-MS/MS of peptides generated from multiple proteases [35] allows multiple oxidation products of ubiquitin to be identified, including several proteoforms of the M+16 Da peak. Under Fe(II)/H2O2 conditions, numerous amino acid modifications are possible [36] and include side chain hydroxylation, carbonylation and backbone cleavage. The variety of these modifications requires multiple database searches be performed [37]. Sample integrity was confirmed by using high resolution ESI-MS on an Orbitrap Velos of intact oxidized protein mixtures.

Experimental Methods

In vitro Oxidation of Ubiquitin

Bovine ubiquitin was purchased from Sigma-Aldrich (St. Louis, MO). Protein (10 mg·mL−1) was dissolved in 10 mM sodium phosphate buffer solution (pH 7.4) and 10 mM H2O2 and 1 mM FeCl2 were added and allowed to react at 37°C for 2 hours. The reaction was quenched by flash freezing with liquid nitrogen. Protein sample was desalted on an HLB cartridge (Waters; Milford, MA) according to manufacturer’s instructions. Solvent was removed by centrifugal evaporation and dried protein stored at −80°C until further analysis.

Top-down ESI-MS and MSn Analysis

Intact oxidized ubiquitin (∼30 μM) was solubilized in 49:49:2 water:methanol:acetic acid. ESI-MS analysis was performed on a LTQ-Orbitrap Velos mass spectrometer (Thermo-Fisher Scientific, Waltham, MA) with direct infusion by a syringe pump. The following electrospray ionization parameters were used: spray voltage 4.25 kV; capillary temperature 200.00°C and flow rate 3 μL·min−1. Orbitrap detector settings included resolving power of 100 k, parent m/z scan range 600–2000, 3 μscans, and 30 and 100 scans for parent and fragmentation spectra, respectively. MS/MS data were recorded in the FT. MS/MS and MS3 settings used an isolation width of 1 m/z and normalized collision energy of 35%.

Protein digestion

Purified oxidized ubiquitin (1 μg·μL−1) was solubilized in a denaturing buffer (0.2 M Tris, 8 M urea, 10 mM CaCl2, pH 8.0). Tris buffer (0.2M Tris, 10mM CaCl2, pH = 8.0) was added to dilute urea to 2M. The solution was separated into three equal volume aliquots and each incubated with TPCK-treated trypsin (Sigma), glutamic acid-C [(Glu-C); Princeton Separation, Inc, Adelphia, NJ] or lysine-C [(Lys-C); Princeton Separation, Inc] proteases at a 1:50 protein:enzyme mass ratio for 24 h at 37°C. Liquid nitrogen was used to quench digestions and samples were acidified by adding formic acid, desalted with HLB cartridges and the eluent dried by centrifugal evaporation.

Nanoflow LC-MS/MS

Online desalting and reversed-phase chromatography was performed with a nanoLC system equipped with an autosampler (Eksigent; Dublin, CA). Mobile phases A and B for these analyses were 96.95:2.95:0.1 water:acetonitrile:formic acid and 99.9:0.1 acetonitrile:formic acid, respectively. Five μL of each peptide sample (1 μg·μL−1 in 0.1% formic acid) was loaded on to a trapping column [100 μm i.d. × 2 cm; 3 μm C18 200 Å stationary phase material (Michrom Bioresource Inc.;Auburn, CA)] at 3 μL·min−1 in 3% mobile phase B for 3 min. After desalting, the sample was loaded onto a pulled-tip (using a CO2 laser) analytical column (75 μm i.d. × 13.2 cm), packed in-house with 3 μm C18 100 Å stationary phase material (Michrom Bioresource Inc.). The following gradient was delivered at a flow rate of 300 nL·min−1: 0–5 min, 10% mobile phase B; 5–15 min, 10–30% B; 15–45 min, 30–45% B; 45–50 min, 45–60% B; 50–55 min, 60–80% B; 55–65 min, 80% B; 65–75 min, 10% B. The LC eluent was introduced into the ESI source with ∼1.5–2.0 kV. Data-dependent acquisition parameters were: parent Orbitrap MS resolving power 60 k; m/z scan range 300–1800; the top eight most intense ions were selected and activated using CID; isolation width 3 m/z; normalized collision energy 35%; dynamic exclusion was enabled with a repeat count of two for a duration of 60 sec; and, a minimum of 5000 ion counts for MS/MS. Samples were analyzed in triplicate.

Data Analysis

Top-down spectra were viewed and analyzed by Xcalibur 2.1 software (Thermo). The Xtract program (Xcalibur) was used to deconvolute the spectra and calculate protein masses. Spectra were manually inspected and the m/z values matched to theoretical b- and y-type ions generated by ProteinProspector v5.9.4 [38]. For peptide data, .RAW files were analyzed with Proteome Discoverer 1.3 software (Thermo) and MS/MS spectra searched against a .fasta file containing the ubiquitin sequence (truncated from the N-terminal region of Uniprot ID P0CH28). Sequest search parameters included two maximum enzyme miscleavages; precursor mass tolerance of 10 ppm; fragment mass tolerance of 0.8 Da; dynamic modifications (see S1 Table) of mono oxidation to Lys, Arg, Pro, Thr, Met, His, Tyr, Ala, Asn, Asp, Glu, Gln, Ile, Leu, Phe, Ser and Val (Ox, +15.995 Da), dioxidation to Lys, Arg, Pro, His, Tyr, Asn, Asp, Phe and Met (DiOx, + 31.990 Da), carbonylation to Lys, Arg, Pro, Glu, Gln, Leu, Ser, Val and Ile (+13.979 Da), deamidation to Gln, Arg and Asn (0.984 Da), decarboxylation to Asp and Glu (-30.010 Da), oxidation of His to Asn (-23.0159 Da) or Asp (-22.032 Da) or aspartylurea (-10.032 Da) or ring open (4.979 Da), carbonylation of Arg to glutamic semialdehyde (GluSA, -43.053 Da), Lys to aminoadipic semialdehyde (AminoAdSA, -1.032 Da) or aminoadipic acid (14.963 Da), Pro to pyrrolidinone (-30.010 Da), and Thr to 2-amino-3-oxo-butanoic acid (Oxd’n, -2.016 Da). Only peptides with medium (p<0.05) and high confidence (p<0.01) as determined from a reverse decoy database search (which in Proteome Discoverer sets appropriate thresholds for XCorr values as a function of charge state) were used for initial filtering of the data [3941]. For final inclusion of peptide hits and localization of modification sites, all MS/MS spectra were manually validated.

Results and Discussion

MS Analysis of Intact Oxidized Ubiquitin

High-resolution ESI-MS analysis of a solution containing only untreated ubiquitin shows an M+16 Da peak that represents <2% of the total unmodified peak intensity in a deconvoluted spectrum (data not shown). Because the untreated sample has very limited sample handling, the M+16 Da species could be from the manufacturing and storage of the protein product, or from the electrospray ionization, (e.g. solution contact with metal needle or the electrolysis of water under high spray voltage. However, utilizing Fenton chemistry, several peaks belonging to oxidized ubiquitin are observed (Fig. 1). The most intense oxidized peak belongs to an M+16 Da species at each charge state measured (i.e., +5-+13). The inset of Fig. 1 shows a zoom-in of the +12 charge state, whereby two protein isotopic distributions are measured for unmodified and oxidized ubiquitin ions. Upon deconvolution of the spectrum it is noted that the M+16 Da species constitutes ~20% of the unmodified abundance. This is a factor of ten increase in M+16 Da ions in comparison to solutions containing only untreated ubiquitin. The masses of the deconvoluted native and M+16 Da peaks are 8564.601 and 8580.592 Da. These values are <5 ppm of the theoretically derived mass values and indicate a mass shift of 15.991 Da. Notably, this shift corresponds to the predominance of monooxygenated proteoforms in the Fe(II)/H2O2-ubiquitin mixture. The incorporation of an oxygen atom does not appear to influence the ionization efficiencies and hence observed signal intensity of ubiquitin in ESI [42], and is directly related to the relative abundance of each proteoform. Other proteoforms are observed in the spectrum: M-114 Da, M-57 Da, M-44 Da, M-16 Da, M+32 Da species, M+48 Da and M+96 Da. However, these peaks are low intensity and it’s possible many proteoforms are not observed in this direct infusion experiment.

thumbnail
Fig 1. Precursor ion mass spectra of oxidized ubiquitin.

In the inset is a zoom-in of the +12 charge state that shows unmodified and oxidized ubiquitin species. The observed mass shift between native and oxidized ubiquitin is indicated in the figure.

https://doi.org/10.1371/journal.pone.0116606.g001

Characterizing Methionine Oxidation Proteoform

Based on the sequence of ubiquitin, we anticipated methionine oxidation. The [M+O+12H]12+ protein peak was isolated and fragmented in the linear ion trap with CID. As shown in Fig. 2 many of the b- type fragment ion peaks are shifted in mass from expected fragments of unmodified ubiquitin ions by 16 Da. Moderate sequence coverage of the intact protein (Fig. 2) was obtained with CID and could be increased using dissociation methods such as ECD and ETD [4346]. Inspection of the lower mass region (m/z 260–410) of the spectrum in Fig. 2 revealed the detection of [b2+O+H]+ and [b3+O+H]+ ions, indicating the oxygen atom addition on residue Met-1 or Gln-2. Based on the higher sensitivity of methionine to oxidation [47], it is very probable that the conversion of the single methionine residue to methionine sulfoxide occurred.

thumbnail
Fig 2. CID MS/MS spectra obtained upon isolation of +12 charge state oxidized ubiquitin species (m/z 716.06, isolation window 1 m/z).

Zoom-in CID MS/MS spectra of the m/z range 260–410. To the right top is the sequence of ubiquitin with observed fragment ions across all z labeled.

https://doi.org/10.1371/journal.pone.0116606.g002

Multiple proteases allow for enhanced sequence coverage in proteomics as some cleavage sites are inaccessible with common enzymes [35] and oxidative modification may also hinder enzymatic cleavage for some residues (e.g., Lys and Arg) [48]. Peptide analysis of oxidized ubiquitin digests using trypsin, Lys-C, and Glu-C proteases is consistent with a M+16 Da methionine sulfoxide proteoform (see Table 1). Several oxidized methionine-containing peptides were identified in nanoLC-MS/MS analyses including doubly-charged mQIFVKTLTGK, mQIFVK, and mQIFVKTLTGKTITLE peptides derived from trypsin, Lys-C, and Glu-C proteases, respectively (Fig. 3). The most predominant peak in the MS/MS spectra of each of these peptides is the doubly-charged precursor ion with the loss of methyl sulfoxide [M+2H-CH3SOH]2+. This precursor ion is consistent with the presence of an oxidized methionine residue and has been observed by others [49,50]. For the spectra in Fig. 3, observed b-ions (including b2 fragments) are shifted by 16 Da which localizes the modification site to Met1. While not performed in these studies, methionine oxidation can also be tested through enzymatic action with peptidyl methionine sulfoxide reductase or by treatment with dithiothreitol [51,52].

thumbnail
Fig 3. CID MS/MS spectra of (a) [mQIFVKTLTGK+2H]2+ with Met1-Ox as observed by trypsin proteolysis, tr = 18.11 min, m/z = 641.36;(b) [mQIFVK+H]2+ with Met1-Ox as observed by Lys-C proteolysis, tr = 21.27 min, m/z = 391.22 and (c) [mQIFVKTLTGKTITLE+2H]2+ with Met1-Ox as observed by Glu-C proteolysis, tr = 26.58 min, m/z = 920.02.

Note that lowercase letters represent the oxidation of methionine to methionine sulfoxide. Ions labeled with asterisks (*) contain modifications.

https://doi.org/10.1371/journal.pone.0116606.g003

thumbnail
Table 1. List of oxidative modifications identified from multiple proteases.

https://doi.org/10.1371/journal.pone.0116606.t001

Mapping Other M+16 Da Proteoforms

Fig. 4A and 4B show MS/MS spectra of the tryptic peptide [LIfAGK+H]+ and the Lys-C peptide [EGIPPDQQRLIfAGK+2H]2+, respectively. Each of these spectra contains fragment ions that localize an oxidative modification to a phenylalanine residue (i.e., Phe45). Phenylalanine contains an aromatic ring which upon oxidation can be modified by hydroxyl radicals at the para-, ortho-, or meta- positions as shown in Fig. 4A [53,54]. Similarly, MS/MS spectra of the tryptic peptide [TLSDyNIQK+2H]2+ and Lys-C peptide [QLEDGRTLSDyNIQK+2H]2+ tentatively assign oxidation of the Tyr59 residue (Fig. 4C and 4D, respectively).

thumbnail
Fig 4. CID MS/MS spectra of (a) [LIfAGK+H]+ with Phe45-Ox as observed by trypsin proteolysis, tr = 22.16 min, m/z = 664.40; (b) [EGIPPDQQRLIfAGK+2H]2+ with Phe45-Ox as observed by Lys-C proteolysis, tr = 22.71, m/z = 842.95; (c) [TLSDyNIQK+2H]2+ with Tyr59-Ox as observed by trypsin proteolysis, tr = 14.33 min, m/z = 549.28 and (d) [QLEDGRTLSDyNIQK+2H]2+ with Tyr59-Ox as observed by Lys-C proteolysis, tr = 22.14 min, m/z = 898.45.

Note that lowercase letters represent the oxidation of phenylalanine and tyrosine. Ions labeled with asterisks (*) contain modifications.

https://doi.org/10.1371/journal.pone.0116606.g004

Overall we observe several peptides which contain a monoxygenated residue (Table 1). Positions of these modifications are: Met1, Phe4, Leu15, Lys33, Phe45, Ser55, Tyr59, Lys63, His68, Leu69, and Leu71. It is possible that each of these peptides arise from different molecules of intact M+16 Da proteoforms. However, it is also likely that they arise from lower intensity M+32 Da or other oxidized proteoforms. Ambiguous identifications include Pro37 which is shifted by 16 Da and could correspond to oxygen incorporation or a carbonyl shift to glutamic semialdehyde.

Identification of Other Oxidative Proteoforms

The most abundant oxidized species that exists in these data arise from the incorporation of oxygen to amino acid side chains however other proteoforms are present. Fig. 5A and 5B are example MS/MS spectra from trypsin and Lys-C peptides [tLSDYNIQK+2H]2+ and [MQIFVkTLTGK+2H]2+, respectively. Fragment ions are present (Fig. 5A) which locate an oxidation site to threonine55. Threonine oxidation results in carbonylation to 2-amino-3-oxo-butanoic acid represented by a mass loss of −2.016 Da. Fig. 5B provides MS/MS fragments which identify Lys6 oxidation represented by a mass loss of 1.032 Da. It is noted that this N-terminal peptide contains an unmodified methionine residue and thus originates from other oxidized proteoforms. Other modifications observed with multiple proteases are provided in Table 1 (and S1 Fig.). It is possible that oxidative modifications present can influence MS/MS fragmentation patterns and this is dependent on particular amino acid and modification type. MS/MS spectra were manually inspected to eliminate ambiguous and low confidence assignments.

thumbnail
Fig 5. CID MS/MS spectra of (a) [tLSDYNIQK+2H]2+ with Thr55-Oxd’n as observed by trypsin proteolysis, tr = 20.28min, m/z = 540.27 and (b) [MQIFVkTLTGK+2H]2+ with Lys6 – AminoAdSA as observed by Lys-C proteolysis, tr = 26.95, m/z = 632.85.

Note that lowercase letters represent the carbonylation of threonine to 2-amino-3-oxo-butanoic acid and lysine to aminoadipic semialdehyde. Ions labeled with asterisks (*) contain modifications.

https://doi.org/10.1371/journal.pone.0116606.g005

Extensive database searches were performed in order to search for many potential oxidative modifications that may occur using metal-catalyzed oxidation. In addition to searching for hydroxylation and carbonylation of Lys, Arg, Thr and Pro residues, our searches also included dioxidation, carbonylation on other amino acids, deamidation, decarboxylation, as well as different oxidative products of His (S1 Table). Examples of new modifications, e.g. ring opening of His68, carbonylation of Ser20, Ile36, Leu69, were successfully identified through these additional database searches (Table 1).

Fe(II)/H2O2 oxidation of ubiquitin leads to multiple M+16 Da and other proteoforms. The most abundant M+16 Da species contains Met1-Ox and is consistent with high oxidation reactivity [55]. M+16 Da and M+32 Da proteoforms were distinguishable using top-down MSn however, CID MS/MS – MSn only provided a limited amount of information about modification sites. Other dissociation methods such as ECD and ETD may provide more extensive sequence coverage and directly localize oxidative modification sites. On the other hand, bottom-up data only provide information about peptides that are observed in solution and do not completely reflect the specific oxidized proteoforms from which they originate. For example, the tryptic peptide mQIFVK could have arisen from an intact M+16 Da ubiquitin molecule or from other oxidized proteoforms. The observation of the tryptic peptide mQIfVK with two oxidized residues and the variety of modified residues listed in Table 1, clearly indicates that many oxidized proteoforms are present.

Thus use of multiple proteases greatly improved our ability to localize oxidative modification sites of oxidized ubiquitin. Specifically, six out of 24 modified sites identified by trypsin are validated by Lys-C and Glu-C experiments (e.g. Met1 is identified by all three proteases). Modifications to residues such as Lys6, Thr12, Glu24 and Ser65 were only identified with Lys-C and Lys33, Glu34 were only identified by Glu-C (See Table 1). We note that incorporation of an enrichment or tagging step may improve the detection of other oxidized proteoforms [2123]. The combination of multiple proteases with iterative database searching was key to identification of so many oxidized sites and proteoforms of ubiquitin. Due to limitations with the number of modifications that can be simultaneously searched with SEQUEST (in Proteome Discoverer), we performed iterative database searches by classifying the modifications into seven groups (methionine oxidation, non-methionine oxidation, carbonylation (+14 Da), deamidation, decarboxylation, histidine oxidation and special carbonylation) [37]. This resulted in each RAW file being searched 22 times. For simple protein mixtures in which one seeks to gain knowledge about the complexity of oxidized proteoforms we present this strategy (combination of top-down mapping, multiple proteases, and iterative database searching) as an alternative to targeted chemistries or enrichment steps.)

Conclusions

This work reports on the detection of oxidized species of Fe(II)/H2O2 oxidized ubiquitin molecules using multiple proteases and iterative database searching of oxidative modifications. Multiple proteases allowed numerous modification sites to be identified and increased confidence in each oxidative site. Multiple proteases also allowed inaccessible sites by trypsin digestion to be available with other proteases. Iterative database searching allowed different types of oxidative modifications to be identified, however this also required manual validation of MS/MS spectra and extended computing times. Under the mM concentrations of oxidizing reagent used, oxidative modifications to ubiquitin included protein carbonylation. The ability to identify distributions of proteoforms for simple systems, such as ubiquitin, using multiple proteases in shotgun proteomics can be extended to larger and more complex protein samples. Complex protein samples will benefit from additional enrichment steps. The utility of multiple proteases combined with iterative database searching of oxidative modifications can be extended to other types of PTMs such as glycosylation, cysteine oxidations, and amidation and has promising applications in pharmaceutical industries for the analysis of intact antibodies.

Supporting Information

S1 Fig. MS/MS spectra of oxidatively-modified peptides identified in Table 1.

https://doi.org/10.1371/journal.pone.0116606.s001

(PDF)

S1 Table. List of the oxidative modifications included in the database search.

https://doi.org/10.1371/journal.pone.0116606.s002

(XLSX)

Author Contributions

Conceived and designed the experiments: RASR LG. Performed the experiments: LG. Analyzed the data: LG RASR. Contributed reagents/materials/analysis tools: RASR. Wrote the paper: RASR LG.

References

  1. 1. Moller IM, Rogowska-Wrzesinska A, Rao RS (2011) Protein carbonylation and metal-catalyzed protein oxidation in a cellular perspective. J Proteomics 74: 2228–2242. pmid:21601020
  2. 2. Finkel T, Holbrook NJ (2000) Oxidants, oxidative stress and the biology of ageing. Nature. 2000/11/23 ed. pp. 239–247.
  3. 3. Levine RL, Stadtman ER (2001) Oxidative modification of proteins during aging. Exp Gerontol. 2001/08/30 ed. pp. 1495–1502.
  4. 4. Madian AG, Regnier FE (2010) Proteomic identification of carbonylated proteins and their oxidation sites. J Proteome Res. 2010/06/05 ed. pp. 3766–3780.
  5. 5. Stadtman ER, Berlett BS (1997) Reactive oxygen-mediated protein oxidation in aging and disease. Chem Res Toxicol. 1997/05/01 ed. pp. 485–494.
  6. 6. Smith LM, Kelleher NL, Consortium for Top Down P (2013) Proteoform: a single term describing protein complexity. Nat Methods 10: 186–187. pmid:23443629
  7. 7. Ntai I, Kim K, Fellers RT, Skinner OS, Smith AD, et al. (2014) Applying label-free quantitation to top down proteomics. Anal Chem 86: 4961–4968. pmid:24807621
  8. 8. Yuan ZF, Arnaudo AM, Garcia BA (2014) Mass spectrometric analysis of histone proteoforms. Annu Rev Anal Chem (Palo Alto Calif) 7: 113–128. pmid:24896311
  9. 9. Peng Y, Gregorich ZR, Valeja SG, Zhang H, Cai W, et al. (2014) Top-down proteomics reveals concerted reductions in myofilament and Z-disc protein phosphorylation after acute myocardial infarction. Mol Cell Proteomics.
  10. 10. Li H, Wongkongkathep P, Van Orden SL, Ogorzalek Loo RR, Loo JA (2014) Revealing Ligand Binding Sites and Quantifying Subunit Variants of Noncovalent Protein Complexes in a Single Native Top-Down FTICR MS Experiment. J Am Soc Mass Spectrom.
  11. 11. Wells JM, McLuckey SA (2005) Collision-induced dissociation (CID) of peptides and proteins. Methods Enzymol 402: 148–185. pmid:16401509
  12. 12. Brodbelt JS, Wilson JJ (2009) Infrared multiphoton dissociation in quadrupole ion traps. Mass Spectrom Rev 28: 390–424. pmid:19294735
  13. 13. Bakhtiar R, Guan Z (2005) Electron capture dissociation mass spectrometry in characterization of post-translational modifications. Biochem Biophys Res Commun 334: 1–8. pmid:15950932
  14. 14. Wiesner J, Premsler T, Sickmann A (2008) Application of electron transfer dissociation (ETD) for the analysis of posttranslational modifications. Proteomics 8: 4466–4483. pmid:18972526
  15. 15. Catherman AD, Durbin KR, Ahlf DR, Early BP, Fellers RT, et al. (2013) Large-scale Top-down Proteomics of the Human Proteome: Membrane Proteins, Mitochondria, and Senescence. Molecular & Cellular Proteomics 12: 3465–3473.
  16. 16. Gregorich ZR, Ge Y (2014) Top-down proteomics in health and disease: challenges and opportunities. Proteomics 14: 1195–1210. pmid:24723472
  17. 17. Dekker L, Wu S, Vanduijn M, Tolic N, Stingl C, et al. (2014) An integrated top-down and bottom-up proteomic approach to characterize the antigen-binding fragment of antibodies. Proteomics 14: 1239–1248. pmid:24634104
  18. 18. Moradian A, Kalli A, Sweredoski MJ, Hess S (2014) The top-down, middle-down, and bottom-up mass spectrometry approaches for characterization of histone variants and their post-translational modifications. Proteomics 14: 489–497. pmid:24339419
  19. 19. Deterding LJ, Bhattacharjee S, Ramirez DC, Mason RP, Tomer KB (2007) Top-Down and Bottom-Up Mass Spectrometric Characterization of Human Myoglobin-Centered Free Radicals Induced by Oxidative Damage. Analytical Chemistry: American Chemical Society. pp. 6236–6248.
  20. 20. Ge Y, Lawhorn BG, ElNaggar M, Sze SK, Begley TP, et al. (2003) Detection of four oxidation sites in viral prolyl-4-hydroxylase by top-down mass spectrometry. Protein Science 12: 2320–2326. pmid:14500890
  21. 21. Lourette N, Smallwood H, Wu S, Robinson E, Squier T, et al. (2010) A top-down LC-FTICR MS-based strategy for characterizing oxidized calmodulin in activated macrophages. Journal of The American Society for Mass Spectrometry 21: 930–939. pmid:20417115
  22. 22. Rogowska-Wrzesinska A, Wojdyla K, Nedic O, Baron CP, Griffiths H (2014) Analysis of protein carbonylation—pitfalls and promise in commonly used methods. Free Radic Res: 1–31.
  23. 23. Fedorova M, Bollineni RC, Hoffmann R (2014) Protein Carbonylation as a Major Hallmark of Oxidative Damage: Update of Analytical Strategies. Mass Spectrometry Reviews 33: 79–97. pmid:23832618
  24. 24. Ciechanover A, Orian A, Schwartz AL (2000) The ubiquitin-mediated proteolytic pathway: mode of action and clinical implications. J Cell Biochem Suppl 34: 40–51. pmid:10762014
  25. 25. Shang F, Taylor A (2011) Ubiquitin-proteasome pathway and cellular responses to oxidative stress. Free Radic Biol Med 51: 5–16. pmid:21530648
  26. 26. Kastle M, Grune T (2011) Protein Oxidative Modification in the Aging Organism and the Role of the Ubiquitin Proteasomal System. Current Pharmaceutical Design 17: 4007–4022. pmid:22188451
  27. 27. Hershko A, Ciechanover A (1998) The ubiquitin system. Annu Rev Biochem 67: 425–479. pmid:9759494
  28. 28. Sharp JS, Becker JM, Hettich RL (2003) Protein surface mapping by chemical oxidation: structural analysis by mass spectrometry. Anal Biochem 313: 216–225. pmid:12605858
  29. 29. Stadtman ER, Oliver CN (1991) Metal-catalyzed oxidation of proteins. Physiological consequences. J Biol Chem 266: 2005–2008. pmid:1989966
  30. 30. Yi D, Perkins PD (2005) Identification of ubiquitin nitration and oxidation using a liquid chromatography/mass selective detector system. J Biomol Tech 16: 364–370. pmid:16522858
  31. 31. Downard KM, Maleknia SD, Akashi S (2012) Impact of limited oxidation on protein ion mobility and structure of importance to footprinting by radical probe mass spectrometry. Rapid Commun Mass Spectrom 26: 226–230. pmid:22223306
  32. 32. McClintock C, Kertesz V, Hettich RL (2008) Development of an electrochemical oxidation method for probing higher order protein structure with mass spectrometry. Anal Chem 80: 3304–3317. pmid:18351783
  33. 33. Aye TT, Low TY, Sze SK (2005) Nanosecond laser-induced photochemical oxidation method for protein surface mapping with mass spectrometry. Anal Chem 77: 5814–5822. pmid:16159110
  34. 34. Shi H, Gu L, Clemmer DE, Robinson RA (2013) Effects of Fe(II)/H2O2 oxidation on ubiquitin conformers measured by ion mobility-mass spectrometry. J Phys Chem B 117: 164–173. pmid:23211023
  35. 35. Swaney DL, Wenger CD, Coon JJ (2010) Value of Using Multiple Proteases for Large-Scale Mass Spectrometry-Based Proteomics. Journal of Proteome Research 9: 1323–1329. pmid:20113005
  36. 36. Xu GH, Chance MR (2007) Hydroxyl radical-mediated modification of proteins as probes for structural proteomics. Chemical Reviews 107: 3514–3543. pmid:17683160
  37. 37. Madian AG, Regnier FE (2010) Profiling carbonylated proteins in human plasma. J Proteome Res 9: 1330–1343. pmid:20121119
  38. 38. Chalkley RJ, Hansen KC, Baldwin MA (2005) Bioinformatic methods to exploit mass spectrometric data for proteomic applications. Methods Enzymol 402: 289–312. pmid:16401513
  39. 39. Borazjani A, Edelmann MJ, Hardin KL, Herring KL, Allen Crow J, et al. (2011) Catabolism of 4-hydroxy-2-trans-nonenal by THP1 monocytes/macrophages and inactivation of carboxylesterases by this lipid electrophile. Chem Biol Interact 194: 1–12. pmid:21878322
  40. 40. Wang L, Cvetkov TL, Chance MR, Moiseenkova-Bell VY (2012) Identification of in vivo disulfide conformation of TRPA1 ion channel. J Biol Chem 287: 6169–6176. pmid:22207754
  41. 41. Robinson RA, Evans AR (2012) Enhanced sample multiplexing for nitrotyrosine-modified proteins using combined precursor isotopic labeling and isobaric tagging. Anal Chem 84: 4677–4686. pmid:22509719
  42. 42. Liu H, Lei QP, Washabaugh M (2014) Evaluation of Label-free MS-based Relative Quantitation of Post-translational Modifications of Therapeutic proteins. ASMS Annual Conference.
  43. 43. Hartmer RG, Kaplan DA, Stoermer C, Lubeck M, Park MA (2009) Data-dependent electron transfer dissociation of large peptides and medium size proteins in a QTOF instrument on a liquid chromatography timescale. Rapid Commun Mass Spectrom 23: 2273–2282. pmid:19575399
  44. 44. Mikesh LM, Ueberheide B, Chi A, Coon JJ, Syka JE, et al. (2006) The utility of ETD mass spectrometry in proteomic analysis. Biochim Biophys Acta 1764: 1811–1822. pmid:17118725
  45. 45. Sze SK, Ge Y, Oh H, McLafferty FW (2002) Top-down mass spectrometry of a 29-kDa protein for characterization of any posttranslational modification to within one residue. Proc Natl Acad Sci U S A. 2002/02/14 ed. pp. 1774–1779.
  46. 46. Sze SK, Ge Y, Oh H, McLafferty FW (2003) Plasma electron capture dissociation for the characterization of large proteins by top down mass spectrometry. Anal Chem 75: 1599–1603. pmid:12705591
  47. 47. Stadtman ER, Levine RL (2003) Free radical-mediated oxidation of free amino acids and amino acid residues in proteins. Amino Acids 25: 207–218. pmid:14661084
  48. 48. Mirzaei H, Regnier F (2005) Affinity Chromatographic Selection of Carbonylated Proteins Followed by Identification of Oxidation Sites Using Tandem Mass Spectrometry. Analytical Chemistry 77: 2386–2392. pmid:15828771
  49. 49. Madian AG, Hindupur J, Hulleman JD, Diaz-Maldonado N, Mishra VR, et al. (2012) Effect of Single Amino Acid Substitution on Oxidative Modifications of the Parkinson's Disease-Related Protein, DJ-1. Mol Cell Proteomics 11: M111 010892. pmid:22104028
  50. 50. Zhao C, Sethuraman M, Clavreul N, Kaur P, Cohen RA, et al. (2006) Detailed Map of Oxidative Post-Translational Modifications of Human P21Ras Using Fourier Transform Mass Spectrometry. Analytical Chemistry 78: 5134–5142. pmid:16841939
  51. 51. Bacarese-Hamilton AJ, Adrian TE, Chohan P, Antony T, Bloom SR (1985) Oxidation/reduction of methionine residues in CCK: a study by radioimmunoassay and isocratic reverse phase high pressure liquid chromatography. Peptides 6: 17–22. pmid:3991360
  52. 52. Hayes CS, Illades-Aguiar B, Casillas-Martinez L, Setlow P (1998) In vitro and in vivo oxidation of methionine residues in small, acid-soluble spore proteins from Bacillus species. J Bacteriol 180: 2694–2700. pmid:9573155
  53. 53. Maskos Z, Rush JD, Koppenol WH (1992) The hydroxylation of phenylalanine and tyrosine: a comparison with salicylate and tryptophan. Arch Biochem Biophys 296: 521–529. pmid:1321588
  54. 54. Solar S (1985) Reaction of Oh with Phenylalanine in Neutral Aqueous-Solution. Radiation Physics and Chemistry 26: 103–108.
  55. 55. Maleknia SD, Brenowitz M, Chance MR (1999) Millisecond radiolytic modification of peptides by synchrotron X-rays identified by mass spectrometry. Anal Chem 71: 3965–3973. pmid:10500483